Skip to main content
Heliyon logoLink to Heliyon
. 2024 Mar 19;10(6):e27761. doi: 10.1016/j.heliyon.2024.e27761

Effect of silver doping on the band gap tuning of tungsten oxide thin films for optoelectronic applications

Md Arifuzzaman a, Tusar Saha a,c, Jiban Podder a,, Fahad Al-Bin b, Hari Narayan Das d
PMCID: PMC10966605  PMID: 38545163

Abstract

In the cutting-edge world, semiconductor metal oxides usually tend to have a high optical band gap (>3.0 eV), significantly acceptable for potential optoelectronic applications. The present study discusses the synthesize of pristine tungsten trioxide (WO3) and Silver (Ag) doped WO3 (Ag: WO3) thin films onto a glass substrate at 450 °C, with varying concentrations of Ag doping (2, 4, 6, 8 and 10 at.%) using a simple Spray Pyrolysis Technique. Field emission scanning electron microscopy (FESEM) analysis showed the presence of particles in the WO3 and Ag: WO3 materials. The X-ray diffraction (XRD) pattern confirmed that the samples' hexagonal structure remained intact. In addition, Rietveld refinement was used for the samples to study the crystal structure meticulously. Because of the surface plasmon resonance effect, the samples' distinguishing characteristics were visible in their optical nature. For pristine WO3, the experimental band gap was determined to be 3.20 eV, and for varying doping concentrations, it was found to be 3.15 eV–2.90 eV, respectively. Furthermore, the fracture has remained imperceptible at elevated concentrations, resulting in a substantial influence on the optical characteristics of 10% Ag: WO3 thin films. The estimated redox potential for 2% Ag: WO3 shows a considerable influence of the band edge potential of the Conduction Band (CB) and Valance Band (VB). The activation energy was determined using temperature-dependent electrical resistivity and exhibited an ohmic nature. The synthesized material exhibited a negative temperature coefficient (NTC) effect at higher concentrations of doping, suggesting its potential applicability as a thermistor. A comprehensive analysis of this present study indicates that Ag can be a viable candidate for doping on WO3 thin films for use in optoelectronic devices.

Keywords: Tungsten trioxide (WO3) thin films, Spray pyrolysis, Structural properties, Band gap energy, Redox potential

Graphical abstract

Image 1

Highlights

  • High lights for review:

  • we reported the influence of Ag doping on the surface morphology, structural, optical and electrical properties of tungsten trioxide (WO3) thin films.

  • The measured band gap for pristine WO3 is found 3.20 eV and then reduced for different concentrations of Ag doping.

  • Narrowing of the band gap and edge of the VB and CB lead to an effective catalyst that could be useful in photo-degradation and a wide range of sensing properties.

  • An effective heat sensing ability is facilitated by negative temperature coefficient in Ag: WO3 thin films.

1. Introduction

The advancement of modern technology is dependent on technical innovation of thin film industry and has a significant impact from the twentieth century for extraordinary performance in solid-state electronics. Thin films' characteristics differ significantly from those of their bulk counterparts. The improvement of optoelectronic devices is the primary goal of the current study, which focuses on creating metal oxide thin films. Tungsten trioxide (WO3) has numerous applications in microelectronics, selective catalysis, environmental engineering, solar cells, gas sensors, photoconductivity, photo-electrochemical cells for solar energy conversion, etc. [[1], [2], [3], [4], [5], [6], [7], [8], [9]]. A comprehensive understanding of the structural property and constancy of WO3 thin films is indispensable for practical device applications. Low-dimensional materials' properties and applications are predominantly determined by their chemical composition, crystal structure, surface morphology and phase stability. The phase transition of WO3 thin films and nanostructures is induced by heating or annealing is essential for specific purposes [10]. Understanding the various phase transitions of WO3 can be attained through an experimental and theoretical approach.

WO3 is an n-type semiconductor metal oxide with a band gap range of approximately 2.7–3.0 eV and visible light irradiation. WO3 has a hexagonal crystal structure with lattice constants of a = 7.298 Å, and c = 3.899 Å and several studies found a monoclinic structure as well [11]. At ambient temperature, the band gap of WO3 is 2.60 eV. Point defects resulting from oxygen vacancies in sub-stoichiometric compounds WO1-x (0<x<1) significantly impact its electronic properties. Because of its large band gap and remarkable electron-transport capability, WO3 is a promising material for usage in a wide variety of optoelectronic devices [12]. In smart windows, large-area information displays, energy-saving, and other applications, films demonstrate remarkable electrochemical stability and are very transparent in the visible spectrum. The analysis of the UV spectrum revealed that, depending on the temperature of the substrate, WO3 has a visible-light transmission of around 70% on average. There is a significant visual transmission at high deposition temperatures [13].

WO3, is a non-toxic transition metal oxide; in addition, the elements that constitute WO3 are cheap and plentiful. Due to interfacial energy, two-dimensional growth is preferred between WO3 and other oxide substrates, leading to high-quality films at reduced temperatures. Sol-gel, electro-spinning, hydrothermal, thermal evaporation, microwave irradiation, spray pyrolysis, sputtering, and electron beam evaporation techniques have been used to prepare WO3 thin films [[14], [15], [16]]. There are benefits and drawbacks to every approach. One of the easiest methods is Spray Pyrolysis Technique (SPT), a non-vacuum system to deposit thin films over large areas with perfectly doped of different transition materials. Diverse transparent metal oxides thin films i.e. SnO2, TiO2, Fe2O3, ZnO have been studied using this technique with different doped materials [[17], [18], [19], [20]]. Moreover, to study or develop new properties, pristine WO3 has been deposited with an alloy of different 3d transition metals [21,22]. Doped WO3 thin films reveal transformed structural, electrical and optical characteristics.

Silver (Ag) is an excellent conductor of both heat and electricity; and has a white color, a soft luster, and a high degree of ductility and malleability. As a result, Ag has been extensively reported to increase the photo catalyst's response to visible light. The ionic radius of Ag (172 p.m.) is lower than that of W (210 p.m.), suggesting the development of small grains. The effective surface area of gas-sensing components increases as grain size decreases [23]. The energy band gap of silver doped nanoparticles shrank with growing Ag content, as shown by absorption spectra. The absorbance band gap in Ag-doped thin films moves to the red as the amount of Ag rises. In pristine WO3, Ag dopant significantly impacts structural, optical, morphological, and electrical properties. The formation of grain size can change due to the incorporation of Ag thin films and finely tune the band gap of variety of oxide materials i.e. WO3, ZnO, etc., [24,25]. However, adjusting the band edge potential can enhance the water-splitting process. Insight into the properties of WO3 composites could lead to new uses for the material with Ag dopant. The present investigation is primarily concerned with the area of optoelectronic applications. This research investigated the synthesis of thin films of WO3 and Ag: WO3 through SPT, with the purpose of using them in optoelectronic devices. To comprehend the effect of Ag content on WO3 thin films, structural, optical, and electrical characterizations were performed on all potential samples and observed the band gap tuning, edge potential, and negative temperature coefficient.

2. Experimental analysis

2.1. Reagents and film preparation

As a precursor of W and Ag sources, reagent grade tungstic acid [H2WO4] (Sigma Aldrich) and silver nitrate [AgNO3] (Merck, Germany) have been used. The SPT has been used to deposit films of pristine WO3 and Ag: WO3. For the synthesis of WO3 thin films, 0.1 M solution of pristine WO3 was produced by mixing distilled water and H2WO4 in a predetermined ratio. A schematic diagram of the experimental set up and the deposition parameters are shown in Fig. 1 and Table 1.

Fig. 1.

Fig. 1

Schematic illustration of a low-cost experimental spray pyrolysis technique set up (Homemade).

Table 1.

Optimize deposition conditions to synthesize pristine WO3 and Ag-doped WO3 thin films by Spray Pyrolysis Technique.

Deposition parameters Optimum value
Temperature 450 ± 5 °C
Concentration 0.1 M
Doping concentration 2.0–10 at. %
Rate of spray 5 mL/min
Air pressure 0.6 bar
Distance between nozzle and substance 25 cm

Then, to dissolve H2WO4, a few drops of ammonia solutions are added to the solution. 2.4985 g of H2WO4 are combined with 100 mL of distilled water to prepare 0.1 M solution. Contrarily, 2, 4, 6, 8, and 10 at.% Ag: WO3 solutions were made by dissolving the calculated amounts of H2WO4 and AgNO3. After 1.5 h of stirring, filtration process has been done by filter paper for separation of suspended solid matter from solution.

2.1.1. Possible reaction

Ammonium Para-tungstate was created when H2WO4 was dissolved in purified water and ammonia solution. Then, WO3 produced through the pyrolysis of ammonium para-tungstate.

Moreover, calculated amount of AgNO3 added for preparing solution of different doped concentration. AgNO3, H2WO4 (precursor), ammonia solution, and water (reactant) combined reaction that may occur on a heated substrate and lead to thin coatings. To produce WO3 thin film on a heated substrate, the following reaction can take place between H2WO4 (precursor), ammonia solution, and water (reactant) [26,27]. As deposited thin films are shown in Fig. 2.

H2WO4+NH4OH+H2O(NH4)10(H2W12O42).4H2O (1)

Then pyrolyzed it,

(NH4)10(H2W12O42).4H2O450±5°CWO3+NH3+H2O (2)
Fig. 2.

Fig. 2

Snapshot of deposited thin films.

Chemical reactions may take place during the formation of doped samples are described below. Distilled water reacts with silver nitrate to generate aqueous silver nitrate.

AgNO3(s)+H2O(l)AgNO3(aq) (3)

AgNO3 decomposes when heated,

AgNO3(aq)Ag+O2+NO2 (4)

The probable chemical reactions of Ag: WO3 are given below,

AgNO3+(NH4)10(H2W12O42).4H2O450±5°CAg:WO3+NH3+H2O (5)

2.2. Characterizations and techniques

The overall surface morphologies and elemental compositions of the as-deposited films were investigated using a FESEM (Model: JEOL JSM-7600F) and an accompanying EDX detector. The UV–visible spectrophotometer (SHIMADZU UV-1601) was employed to record the UV–visible NIR transmission bands at room temperature. In addition, the Fizeau fringes method is used to observe the thickness of films. Likewise, the Four Point Probe method has been utilized to evaluate electrical properties. The Rietveld refinement was carried out with the Fullprof Suit software. All the graphs were plotted using the Origin software. The ImageJ software was utilized to ascertain the magnitude of the crack's thickness on the film's surface.

3. Result and discussions

3.1. Structural analysis

The XRD technique was employed to perform a structural investigation of Ag: WO3 samples, where 0, 2, 4, 6, 8 and 10% Ag concentrations were examined. Fig. 3 displays the outcomes of the Rietveld refinement conducted on samples of Ag: WO3. The pristine WO3 exhibits diffraction peaks at 2 θ values of 13.98°, 23.32°, 24.30°, 27.28°, 28.12°, 33.58°, 36.86°, 49.78°, and 55.67° which are indexed to the corresponding diffractions from (100), (001), (110), (101), (200), (111), (201), (220) and (221) planes, respectively [28,29]. The curve fitting analysis reveals the presence of symmetry in the P6/mmm space group, and the XRD data was analyzed using the Rietveld refinement approach [30]. Hence, the hexagonal structure was confirmed from the XRD and Rietveld refinement. Fig. 3 (a) represent the hexagonal crystal structure of WO3. The figure displays the analysis of the XRD data, with the experimental data (Iobs) represented by black circles and the estimated intensities (Ical) shown by red lines. The structural parameters acquired for the examined samples are displayed in Table 2. The disparity in ionic radii between the parent ion (W6+) and the dopant ion (Ag+) leads to a relatively small displacement of the XRD peaks towards higher 2θ values. As a result, as Ag doping concentrations increased, lattice parameters experienced slight contraction, except for 10% Ag: WO3. The Debye-Scherer formula was employed to determine the average crystallite size of all samples by analyzing the reflected peaks in the X-ray diffraction patterns. It was assumed that the crystallite size is correlated with peak broadening. The estimated average crystallite size of all samples found was within the range of 25–28 nm. The size of crystallites and lattice strain have an impact on X-ray diffraction patterns. The W–H analysis is employed to differentiate between the deformation peak caused by the size and strain of the crystallite, considering the widening of the peak width as a function of 2θ. Scherrer method is most popular method for crystallite size estimation. However, this method is confined with limitation whether intrinsic strain is developed due to point defect, grain boundary triple junction and stacking faults [31,32]. Therefore, the Williamsons Hall method (W–H) and modified W–H method have been implemented here that includes the intrinsic strain along with the crystallite size [[33], [34], [35], [36], [37]]. It implies that the hexagonal structure of the WO3 thin film is unaffected by Ag ions, and it cannot affect the crystalline phase of WO3 [38].

Fig. 3.

Fig. 3

Fig. 3

Rietveld refinement plots illustrating the composition of pristine WO3 and thin films containing 2, 4, 6, 8 and 10 at. % of Ag: WO3.

Fig. 3 (a): Polyhedral hexagonal structure of WO3 found from Rietveld refinement.

Table 2.

Calculated data from XRD analysis.

Sample Lattice Parameters
c/a V (Å3) Average Crystallite Size, D (nm) W–H Method UDM
χ2
a = b (Å) c (Å) D (nm) ε (10−3)
Pristine WO3 7.32 3.81 0.52 176.80 27 46 1.8 4.1
2 at.% Ag:WO3 7.31 3.81 0.52 176.31 26 41 1.47 5.5
4 at.% Ag:WO3 7.30 3.81 0.52 175.83 28 72 2.93 6.3
6 at.% Ag:WO3 7.30 3.81 0.52 175.83 25 69 3.09 6.6
8 at.% Ag:WO3 7.30 3.81 0.52 176.31 28 71 2.68 5.8
10 at.% Ag:WO3 7.32 3.81 0.52 176.80 26 58 2.68 4.8

According to this model, the total broadening of XRD peak (βtotal) is assumed as, βtotal=βhkl+βstrain. In the following sections the structural analysis by both W–H and modified W–H method have been discussed and average crystallite size (D) and microstrain (ε) have been estimated. The uniform deformation model (UDM) considers the strain that arises in nanostructures of a defective crystal due to its isotropic nature [35]. Here, physical broadening of the XRD profile due to intrinsic strain (βstrain) has been considered as, βstrain=4ε.tanθ where θ is the Bragg position,

β=0.9λD×1cos(θ)+4ε.tanθ (6)

Now the further rearrangement of equation [6] is as,

β*cos(θ)=0.9λD+4ε.sinθ (7)

Now, the equation [7] has been represented graphically (Fig. 4) with the term 4sinθ as independent variable (along X-axis) and β*cos(θ) as a dependent variable (along Y-axis). This graph is a representation of the UDM model. The slope of linear fitting of this will provide the value of intrinsic strain and intercept gives average crystallite size (D).

Fig. 4.

Fig. 4

(UDM) The variation of βCos(θ) on the Y-axis and 4sin(θ) on the X-axis for the set of reflection data for different samples.

The lattice strain arises from the alteration of the lattice volume, which may either expand or shrink. This alteration is limited by the quantum size confinement that governs the behavior of the nanostructures. Consequently, the atomic configuration will undergo a modest modification in comparison to its original structure. Furthermore, the imposition of size restrictions will give rise to many imperfections within the lattice structure, thereby inducing lattice strain. The average crystallite size, as determined by the W–H model, lay within the range of [41–72] nm. This value is considerably larger than the size estimated by the Debye-Scherer formula, which incorporated the term of intrinsic strain to obtain the W–H model.

The original crystal, that is free of Ag has a certain value of intrinsic strain which reflects the effects of impurities and temperature. Introducing Ag to the crystal causes an increase in intrinsic strain values, these increases go proportionally for 4 and 6 at. % Ag dope WO3. But it decreases slightly for 8 and 10 at. % Ag doped WO3, which indicates a kind of relaxation inside the crystal. The bigger Ag+ (radius = 1.26 Å) ions absorb the W6+ (radius = 0.64 Å) ions inside the crystal when Ag concentrations rise. Hence, this process aims to attain the relaxation state by reducing the internal energy and total entropy of the crystal by relocating all the atoms throughout the whole gaps. The subsequent phase will include an increase in inherent strain due to an increase in the Ag concentration. Furthermore, this phenomenon also demonstrates the stochastic fluctuations in crystal size that occur when different concentrations of Ag are introduced, as the crystal attempts to optimize its energy and entropy by rearranging the locations of its atoms within the given space. Table 2 presents data on the intrinsic strain, lattice parameters, chi square, c/a ratio, and crystallite size ascertained using the Debye-Scherer formula and the W–H model.

The quantitative determination of the feature of the WO3 and Ag: WO3 thin film may be obtained using the texture coefficient (TChkl). This coefficient is estimated based on the X-ray diffraction (XRD) analysis of the (hkl) plane, using the following formula [17]:

TChkl=I(hkl)I0(hkl)1NI(hkl)I0(hkl)

The equation relates the measured intensity (I) of the WO3 film to the standard intensity (I0) and the number of (hkl) diffraction peaks (N). An analysis of a randomly oriented crystallite shows a TChkl value of 1. If TChkl is more than 1, it indicates a higher prevalence of crystallite orientation. For varying concentrations of Ag, Fig. 3 shows the corresponding Texture Coefficient values for the three main diffraction planes (100), (001), and (200), and Table 3 lists these values. The TC value for the (100), (001), and (200) planes exhibit noticeable variations with increasing Ag-doping concentrations. These findings suggest a little re-orientation impact due to the Fe-doping ratio, which might explain the reduction in cracking thickness.

Table 3.

Texture coefficient of WO3 and Ag: WO3 samples.

Samples Orientation Plane (hkl) Texture coefficient TChkl
WO3 100 1.00
001 1.16
200 0.80
2% Ag: WO3 100 1.13
001 0.97
200 1.60
4% Ag: WO3 100 0.97
001 1.04
200 1.17
6% Ag: WO3 100 1.46
001 0.99
200 0.92
8% Ag: WO3 100 1.04
001 1.08
200 0.79
10% Ag: WO3 100 1.14
001 1.09
200 0.76

3.2. Surface analysis

FESEM images of pristine WO3 and Ag: WO3 thin films at 3k magnification and 5.0 kV are shown in Fig. 5. These coatings are deposited on substrates heated to 450 °C. The synthesis conditions and surface chemistry of WO3 thin coatings can significantly affect their shape stability. The Ag: WO3 FESEM micrograph shows granules with distinctive spherical shapes. Considering this, it is permissible for thin coatings of WO3 to incorporate atoms of silver. Fig. 5. (a, b, c, and d) illustrate surface micro cracks. It has been observed that significant influence of thickness on fracture formation [39]. The widths of the cracks in diameter are shown in Table 4. However, no crack is observed for the higher percentage of Ag doping (especially 10% Ag doping). Merely cracks are found for 8% Ag doping. In Fig. 6 depicts the zoom view of crack formation of all samples except 10% doped Ag: WO3 thin film.

Fig. 5.

Fig. 5

(a–f) FESEM images of pristine WO3 and 2, 4, 6, 8 and 10 at. % Ag: WO3 thin films at 3k magnifications.

Table 4.

Average particle size and crack width for each as-deposited thin film.

Sample Average particle size in diameter, (μm) Crack width, (μm) Standard deviation
Pristine WO3 3.40 0.33 0.10
2% Ag 3.41 0.11 0.04
4% Ag 3.62 0.10 0.02
6% Ag 3.55 0.15 0.06
8% Ag 3.79 0.07 0.01
10% Ag 1.85

Fig. 6.

Fig. 6

(a–e) Zoom view of FESEM images for pristine WO3 and 2 to 8 at. % Ag: WO3 thin films.

The addition of Ag produces sphere grains and reduces the crack's thickness. These fractures result from an imbalance between the compressive stress, thermal expansion of WO3 thin coatings and the glass substrate. Introducing Ag by doping can incorporate impurity atoms into the crystal lattice of the film material. Impurity atoms can immobilize dislocations, which are imperfections in the arrangement of atoms in a crystal structure that can serve as starting points for cracks. Ag doping can significantly decrease the mobility of dislocations by immobilizing them, hence reducing the propagation and growth of fractures. According to our findings, when the concentration of Ag doping is increased, cracks would become invisible.

Ag dopant is incorporated into WO3 thin films as micro-particles of Ag and W, combined on the surface of Ag: WO3 thin films. The particle size increased as Ag concentration increased. Image of 10 at.% Ag: WO3 film reveals the formation of incompatible structure grains and spheres. It is likely because a higher concentration of Ag could disrupt the crystal matrix and impede the thin film's growth.

The higher the concentration of Ag (10 at.%), the grains become the smaller due to Ag's collision with W. In addition, increasing particle size leads to increased absorbance [40]. As the particle size decreases, the absorbance of 10 at.% Ag-doped material has shown the decline nature. For particles in a fluid, such as powder or granular material, the GSD is an ensemble of numbers or mathematical functions that characterizes the distribution of particle sizes. Fig. 7 depicts the histogram of the GSD for all synthesized samples. From FESEM pictures, ImageJ software was used to reveal a histogram for GSD.

Fig. 7.

Fig. 7

Histogram of the Gaussian size distribution (GSD) of pristine WO3 and Ag: WO3 thin films.

3.3. Elemental analysis

For the as-deposited film, the EDX method was used for the quantitative examination of the films. Fig. 8 shows the results of an EDX analysis performed on all samples, confirming the existence of W and O.

Fig. 8.

Fig. 8

EDX spectrum of thin films made of (a) pristine WO3, (b) 2, (c) 4, (d) 6, (e) 8, and (f) 10 at.% Ag: WO3.

These elements can be seen as two distinct peaks, with respective mass percentages of 74.1 and 25.9. Moreover, the spectrum reveals no impurity originating from carbon additives. Ghasemi et al. [41] reported an equivalent EDX analysis outcome. In thin films of pristine WO3, the proportion of W is significantly greater than that of O. The EDX Fig. 8 for doping of Ag confirmed the presence of Ag at the 3.00 eV position. The mass percentages (mass%) of W, O and Ag are precisely understandable through the graphical representation in Fig. 9.

Fig. 9.

Fig. 9

Mass percentages of every element of deposited sample for different Ag concentration.

3.4. Optical properties

Data from a UV–Vis absorption spectrometer has been used to investigate the samples' optical characteristics. This absorbance graph showed that the samples have slight variations from one another. Absorbance changes significantly when Ag dopants are introduced.

3.4.1. Absorption coefficient analysis

The optical absorption coefficient of all samples are shown in Fig. 10 using UV–Vis spectra. In Fig. 10 without 10 at.% Ag doping, the absorption coefficient of the samples rises with doping concentrations of Ag. Surprisingly, the absorption coefficient of WO3 thin films doped with 10 at.% Ag is lower than films doped with 6 and 8 at.% Ag. When measured in wavelength units, absorption coefficient begins at 360 nm. As the wavelength rises, both films become less absorption coefficient. This drop in absorption coefficient is characteristic of a material having an optical band gap [42]. Therefore, Ag: WO3 thin films have higher absorption coefficient in the visible region than their pristine counterparts. WO3 thin films with higher sensitivities to visible light due to Ag contamination may be susceptible to photo-degradation, according to this evidence. Since absorption coefficient declines with increasing silver incorporation, the transparency of the films gradually decreases. There are several potential causes for the decrease in transmittance that accompanies the increased incorporation of silver. The presence of Ag nanoparticles may cause a reduction in transmittance due to Surface Plasmon Resonance (SPR). A phenomenon known as SPR occurs when the collective oscillation of CB electrons occurs in resonance with the oscillating electric field of incoming light, causing the non-radiative excitation of strong plasmonic electrons. Ag inclusion causes this behavior. A few other materials, like gold and aluminum, are also responsible for this phenomenon. This phenomenon has been reported by several publications as a consequence of Ag inclusion in WO3 thin films [43,44].

Fig. 10.

Fig. 10

Absorption coefficient of pristine and Ag: WO3 thin films for different wavelengths.

3.4.2. Optical band gap

The energy difference between the bands has been calculated using UV–Vis Spectrometry. Optical absorbance has been used with the Tauc equation to determine the optical energy band gap of the films.

αhν=A(hνEg)n (8)

Here, α represent the absorption coefficient, ℎν depicts the photon energy, n is the parameter connected to the distribution of the density of states, and A is a constant or Tauc parameter [15].

The film thickness (t) is determined by measuring the displacement (h) of the fringe system over the film substrate step and using the relation [45],

t=hfringsspacing×λ2

where λ is the used monochromatic light's wavelength. We may write as follows if the fringe spacing is (l),

t=hλ2l

Fig. 11 indicates the linear band gap plots for (αℎν)2 vs E (eV). Here, the direct nature n is allowed to be 1/2, and the indirect nature n is allowed to be 2. An electron that is located in the CB of a direct nature band gap semiconductor has the ability to travel to an empty state in the VB, thereby releasing the energy difference. In comparison, an electron in the lowest part of the CB of an indirect band gap semiconductor cannot go straight down into the highest part of the VB. But along with its energy, it must also change its momentum. The linearity of these plots indicates the transitions in these films are indirect. Using the intersection of the plot's tangent and the plot, the band gap value has been determined. The literature and the findings we reported are consistent regarding the indirect band gap of 3.2 eV for pristine WO3 thin films [12,43]. In the range of 0–8 at.% Ag doping concentration, the band gap of a thin WO3 film decreases from 3.2 to 2.9 eV, respectively. The calculated value of band gap and crystallite size of pristine and WO3 with varying amounts of Ag dopant are shown graphically in Fig. 12. Table 5 represents the calculated band gap and thickness of the all samples.

Fig. 11.

Fig. 11

Tauc's plot of (αℎν)2 vs E (eV) of pristine and doped WO3 thin films with different Ag concentrations.

Fig. 12.

Fig. 12

Graphical representation of optical band gap and crystallite size for various Ag concentrations on WO3 thin films.

Table 5.

Bandgap and thickness of all synthesized thin films.

Sample Bandgap, (eV) Thickness of the sample, (nm)
Pristine WO3 3.20 635 ± 20
2% Ag 3.15 631 ± 20
4% Ag 3.07 638 ± 20
6% Ag 2.95 627 ± 20
8% Ag 2.90 636 ± 20
10% Ag 3.02 641 ± 20

Narrowing the band gap and visible light absorption could be happened due to the oxygen vacancies, and the following interpretation of these data is also possible: Ag-doping creates new localized energy states between tungsten's valence and conduction bands at low at% of doping (up to 8%). These localized energy states function as newly introduced trapping levels, which cause Ag-doped tungsten trioxide thin films' bandgap to decrease. For 10 at.% Ag: WO3 thin film has a band gap value of 3.02 eV. This change, a red shift in Fig. 10 aligns with absorption data. The results have shown that adding Ag to WO3 thin films could have an essential effect on reducing the impact of the band gap energy. Several past studies agree with this observation of decreasing band gap energy with indirect bandgap [20,46,47].

3.5. Redox potential

The charge carrier generation and redox potential are crucial for understanding the semiconductor materials' water splitting (Photo-catalytic Activity). The position of CBM (Conduction Band Maximum) VBM (Valance Band Minimum) of the synthesized pristine and Ag: WO3 thin films are determined for studying the redox potential for this present study.

The semiconductor band-edge locations are key factors for identifying the energetics of PEC (Photo Electrode Chemical) water-splitting devices [48]. The following equations are utilized to compute the band edge of VB and CB of our synthesized semiconductor catalysts [49].:

ECB=XEe12Eg (9)
EVB=ECB+Eg (10)

Here, Eg, the calculated energy of band gap is calculated from the plot (using equation (10)), where ECB and EVB are the band edge potentials of CB and VB, respectively, X represents the calculated Mulliken electro-negativity, Ee is the energy of the free electron on the hydrogen balance (4.5 eV), and so on. In light of the above equation, Fig. 13 illustrates the computed ECB/ EVB values of pristine and Ag: WO3 films and the potential band gap concerning the NHE (Normal Hydrogen Electrode) (pH = 0). Given that the VB edge potential (O2/H2O) is 1.23 eV and the CB edge potential (H+/H2) is 0 eV, the theoretically lowest band gap for water electrolysis is 1.23 eV. Different Ag doping concentrations on WO3 thin films cause the edge of the CB and valence band to change. The downward movement of the CB's leading edge brings it closer to the H+/H2 reduction potential.

Fig: 13.

Fig: 13

Schematic diagram of positions of CBM and VCM of pristine and different concentration of Ag: WO3 thin films concerning the normal hydrogen electrode (NHE).

In Ag2O/WO3 crystals, incident photon energies (ℎν) are absorbed by the n-type WO3 semiconductor, which then excites electrons from the WO3 VB to the WO3 CB and causes them to migrate to the p-type Ag2O semiconductor, facilitating electron-hole separation. For a semiconductor photo-catalyst, edge of valence band (VB) must have positive than oxidation potential of O2/H2O. In contrast, the CB edge position must have less than zero [50]. In this study, Ag dopant has been effective for tuning the band edge potential and enhancing the photo degradation process of the WO3 thin films.

When the semiconductor photo-catalyst is exposed to radiation (hʋ), the hole (h+) and electron (e) are created for jumping the electron VB to CB. Those hole and electron are responsible for forming free radicals, and the free radicals (hydroxyl) are oxidative. For producing free radicals spontaneous photo-degradation process is occurred.

3.6. Electrical properties

3.6.1. I–V curve

I–V characteristic curve also known as current-voltage curves describing how an electrical device or component operates in an electrical circuit. The voltages used to measure the current and voltage drops are examined in the range of 10–65 V. For various voltages, the current passing through pristine WO3 thin films is less than that passing through Ag: WO3 thin films. Hence, Ag is a highly conductive material, adding Ag to WO3 thin films causes the current flow to rise.

I–V curves are practically linear and show the ohmic nature, as depicted in Fig. 14. The results indicate that the rising doping concentrations of Ag ions enhance the current flow. This ohmic behavior of WO3 thin films is comparable to prior studies [51].

Fig. 14.

Fig. 14

I–V Characteristic curves for pristine WO3 and Ag: WO3 thin films.

3.6.2. Variation of resistivity with temperature

Ag doping substantially reduces the resistivity of pristine WO3 thin films at room temperature, which is 19.63×104 Ω-cm. Resistivity results from a trade-off between two simultaneously completing processes [52]. It is believed that the hopping conduction process controls the electronic transport of WO3, and electrons are the primary carriers of oxygen vacancies. Fig. 15 has depicted the temperature dependence electrical resistivity of each sample. Temperature-dependent decreases in resistivity reveal the behavior of semiconductors is the Negative Temperature Coefficient (NTC). The semiconductors with NCT utilize it to understand the signal for temperature changes.

Fig. 15.

Fig. 15

Temperature vs resistivity curves of pristine and Ag: WO3 thin films.

3.6.3. Activation energy

The four-point probe method was used to characterize the temperature dependence conductivity of pristine and Ag: WO3. This work has shown that pristine and Ag-doped samples affect electrical properties significantly. The following Arrhenius law was applied,

σ=σ0exp(EactkT) (11)

In the following equation, σ represents conductivity, σ0 represents the pre-exponential factor, Eact represents the activation energy for conduction, k represents the Boltzmann constant, and T represents the absolute temperature.

Equation [13] can be expressed as,

Inσ=(ΔEKBT)+Inσ0 (12)

The slopes of lnσ vs 1000T have plotted using the Arrhenius relation to determine the activation energy. A plot of lnσ vs 1000T is shown in Fig. 16. From the resistivity curve, it has been confirmed that higher temperatures result in higher conductivity; similarly, the activation energy graph as well [53]. The calculated value of activation energy is enlisted in Table 6.

Fig. 16.

Fig. 16

The changes in activation energy for pristine and Ag: WO3 thin films as the temperature varied.

Table 6.

Calculated value of activation energy.

Sample Activation energy, Δ E1 (eV) (308K–348K) Activation energy, Δ E2 (eV) (353K–393K)
Pristine WO3 0.74 0.89
2% Ag:WO3 0.68 0.92
4% Ag:WO3 0.80 0.89
6% Ag:WO3 0.67 0.83
8% Ag:WO3 0.58 0.78
10% Ag:WO3 0.64 0.71

Thus, the activation energy ΔE can be represented by the following equation,

ΔE=lnσ1TKB (13)

When reactant molecules are heated, their mobility increases, leading to more frequent and forceful collisions taking place within individual molecules between the atoms and bonds. This increases the likelihood of bond rupture, which is necessary for the chemical reaction to occur. For bonds to be broken, the molecule must enter the unstable transition state. This state requires activation energy to be added to the molecule [54]. However, because the transition state is unstable, reactant molecules quickly move onto the next stage of the reaction rather than remaining in the transition state for long periods.

4. Conclusions

A simple spray pyrolysis method was used to deposit WO3 thin films onto glass substrates and revealed both pristine and Ag: WO3 films are to be polycrystalline with a hexagonal crystal structure. The addition of Ag atoms leads to an inconsistent calculation of crystallite size using both the Debye-Scherer formula and the W–H plot. The crystal lattice defect is responsible for the maximum lattice strain seen in WO3 doped with 6 at. % of Ag. Thin films of pristine WO3 grew in the (200) direction while doping with Ag caused a small shift in the select orientation plane. Different amounts of Ag doping were found to significantly alter the surface morphology of WO3, as observed by FESEM. The absorption spectrum was consistently affected by the particle size distribution. UV–Vis spectroscopy revealed a red shift and a reduction in the width of the optical band gap, which is particularly significant for optoelectronic applications. Furthermore, the adjustment of conduction and valence band edges by Ag doping has a substantial impact on the photocatalytic process and 2 at. % Ag-doped WO3 exhibits notable performance in the photocatalytic process. Both pristine and Ag: WO3 thin films exhibit a negative temperature coefficient behaviour, making them suitable for use as heat sensors. Based on a thorough review of this work, it is evident that Ag has the potential to be an appropriate candidate for doping WO3 thin films to enhance their performance in optoelectronic devices.

Funding

This research did not receive any specific grant from funding agencies in the public, commercial, or not-for-profit sectors.

Data statement

The data used for this article is derived from the experiments. These are not available online.

CRediT authorship contribution statement

Md Arifuzzaman: Writing – original draft, Validation, Investigation, Formal analysis, Data curation. Tusar Saha: Validation, Methodology, Investigation, Formal analysis, Data curation. Jiban Podder: Writing – review & editing, Validation, Supervision, Methodology, Investigation, Data curation, Conceptualization. Fahad Al-Bin: Validation, Formal analysis, Data curation. Hari Narayan Das: Validation, Formal analysis, Data curation.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Acknowledgments

This complete research work is carried out in the Spray Pyrolysis lab of the Department of Physics, Bangladesh University of Engineering and Technology (BUET), Dhaka, Bangladesh. The authors are thankful to the Atomic Energy Centre, Dhaka, Bangladesh Atomic energy Commission for using the facilities of FESEM, EDX, XRD, UV–visible spectrometer for characterization of the as deposited thin films.

References

  • 1.Bertus L.M., Enesca A., Duta A. Influence of spray pyrolysis deposition parameters on the optoelectronic properties of WO3 thin films. Thin Solid Films. 2012;520:4282–4290. doi: 10.1016/j.tsf.2012.02.052. [DOI] [Google Scholar]
  • 2.Lemire C., Lollman D.B., Al Mohammad A., Gillet E., Aguir K. Reactive RF magnetron sputtering deposition of WO3 thin films. Sens. Actuators B: Chem. 2002;84:43–48. doi: 10.1016/S0925-4005(02)00009-6. [DOI] [Google Scholar]
  • 3.Chandrasekaran G., Sundararaj A., Therese H.A., Jeganathan K. Ni-catalysed WO3 nanostructures grown by electron beam rapid thermal annealing for NO2 gas sensing. J. Nanoparticle Res. 2015;17:1–11. doi: 10.1007/s11051-015-3100-8. [DOI] [Google Scholar]
  • 4.Mi Q., Ping Y., Li Y., Cao B., Brunschwig B.S., Khalifah P.G., Galli G.A., Gray H.B., Lewis N.S. Thermally stable N2-Intercalated WO3 photoanodes for water oxidation. J. Am. Chem. Soc. 2012;134:18318–18324. doi: 10.1021/ja3067622. [DOI] [PubMed] [Google Scholar]
  • 5.Bathe S.R., Patil P.S. Electrochromic characteristics of fibrous reticulated WO3 thin films prepared by pulsed spray pyrolysis technique. Sol. Energy Mater. Sol. Cells. 2007;91:1097–1101. doi: 10.1016/j.solmat.2007.03.005. [DOI] [Google Scholar]
  • 6.Wang X., Yu C., Wu J., Zhang Y. Fabrication of a highly sensitive and selective Ag-doped WO3 formaldehyde gas sensor. Chem. Lett. 2012;41:595–596. doi: 10.1246/cl.2012.595. [DOI] [Google Scholar]
  • 7.Hao J., Studenikin S.A., Cocivera M. Transient photoconductivity properties of tungsten oxide thin films prepared by spray pyrolysis. J. Appl. Phys. 2011;90:5064–5069. doi: 10.1063/1.1412567. [DOI] [Google Scholar]
  • 8.Simchi H., McCandless B.E., Meng T., Shafarman W.N. Structural, optical, and surface properties of WO3 thin films for solar cells. J. Alloys Compd. 2014;617:609–615. doi: 10.1016/j.jallcom.2014.08.047. [DOI] [Google Scholar]
  • 9.Bussolotti F., Lozzi L., Passacantando M., Rosa S., Santucci S., Ottaviano L. Surface electronic properties of polycrystalline WO3 thin films. Surf. Sci. 2003;538:113–123. doi: 10.1016/S0039-6028(03)00696-4. [DOI] [Google Scholar]
  • 10.Ramana C.V., Utsunomiya S., Ewing R.C., Julien C.M., Becker U. Structural stability and phase transitions in WO3 thin films. J. Phys. Chem. B. 2006;110:10430–10435. doi: 10.1021/jp056664i. [DOI] [PubMed] [Google Scholar]
  • 11.Harshulkhan S., Janaki K., Velraj G., Sakthi Ganapathy R., Krishnaraj S. Structural and optical properties of Ag doped tungsten oxide (WO3) by microwave-assisted chemical route. J. Mater. Sci. Mater. Electron. 2016;27:3158–3163. doi: 10.1007/s10854-015-4138-1. [DOI] [Google Scholar]
  • 12.Ashtiani H., Bahari A., Gholipour S., Hoseinzadeh S. Structural, optical and electrical properties of WO3–Ag nanocomposites for the electro-optical devices. Appl. Phys. Mater. Sci. Process. 2018;124:12. doi: 10.1007/s00339-017-1412-5. [DOI] [Google Scholar]
  • 13.Shanmugasundaram K., Thirunavukkarasu P., Ramamurthy M., Balaji M., Chandrasekaran J. Growth and characterization of jet nebulizer spray deposited n-type WO3 thin films for junction diode application, Orient. J. Chem. 2017;33:2484–2491. doi: 10.13005/ojc/330542. [DOI] [Google Scholar]
  • 14.Lee K.D. Deposition of WO3 thin films by the sol-gel method. Thin Solid Films. 1997;302:84–88. doi: 10.1016/S0040-6090(96)09556-9. [DOI] [Google Scholar]
  • 15.Khan M.M., Kumar S., Ahamad T., Alhazaa A.N. Enhancement of photocatalytic and electrochemical properties of hydrothermally synthesized WO3 nanoparticles via Ag loading. J. Alloys Compd. 2018;743:485–493. doi: 10.1016/j.jallcom.2018.01.343. [DOI] [Google Scholar]
  • 16.Boonma P., Jaroenapibal P., Horprathum M., Pornnimitra S., Charoen B., Triroj N. Impedance spectroscopic inspection toward sensitivity enhancement of Ag-doped WO3 nanofiber-based carbon monoxide gas sensor. Mater. Sci. Forum. 2016;872:230–234. doi: 10.4028/www.scientific.net/MSF.872.230. [DOI] [Google Scholar]
  • 17.Babu M.H., Dev B.C., Podder J. Texture coefficient and band gap tailoring of Fe-doped SnO2 nanoparticles via thermal spray pyrolysis. Rare Met. 2019:1–10. doi: 10.1007/s12598-019-01278-3. [DOI] [Google Scholar]
  • 18.Sharmin M., Jiban P. Band gap tuning, n-type to p-type transition and ferrimagnetic properties of Mg doped α-Fe2O3 nanostructured thin films. J. Alloys Compd. 2020;818 doi: 10.1016/j.jallcom.2019.152850. [DOI] [Google Scholar]
  • 19.Saha T., Podder J., Islam M.R., Das H.N. Effect of tungsten doping on the microstructure, optical and photocatalytic activity of titanium dioxide thin films deposited by spray pyrolysis. Opt. Mater. 2022;133 doi: 10.1016/j.optmat.2022.113065. [DOI] [Google Scholar]
  • 20.Badawi A., Althobaiti M.G., Ali E.E., Alharthi S.S., Alharbi A.N. A comparative study of the structural and optical properties of transition metals (M= Fe, Co, Mn, Ni) doped ZnO films deposited by spray-pyrolysis technique for optoelectronic applications. Opt. Mater. 2022;124 doi: 10.1016/j.optmat.2022.112055. [DOI] [Google Scholar]
  • 21.Acosta D.R., Pal U. Electrochimica Acta Improving electrochromic behavior of spray pyrolised WO3 thin solid films by Mo doping. Electrochim. Acta. 2011;56:2599–2605. doi: 10.1016/j.electacta.2010.11.038. [DOI] [Google Scholar]
  • 22.Gaury J., Kelder E.M., Bychkov E., Biskos G. Characterization of Nb-doped WO3 thin films produced by electrostatic spray deposition. Thin Solid Films. 2013;534:32–39. doi: 10.1016/j.tsf.2013.01.080. [DOI] [Google Scholar]
  • 23.Jaroenapibal P., Boonma P., Saksilaporn N., Horprathum M., Triroj N. Improved NO2 sensing performance of electrospun WO3 nanofibers with silver doping. Sens. Actuators B Chem. 2018;255:1831–1840. doi: 10.1016/j.snb.2017.08.199. [DOI] [Google Scholar]
  • 24.Adilakshmi G., Reddy R.S., Reddy A.S., Reddy P.S., Reddy C.S. Ag-doped WO3 nanostructure films for organic volatile gas sensor application. J. Mater. Sci. Mater. Electron. 2020;31:12158–12168. doi: 10.1007/s10854-020-03762-4. [DOI] [Google Scholar]
  • 25.Althobaiti M.G., Alharthi S.S., Alharbi A.N., Badawi A. Impact of silver/copper dual-doping on the structure, linear and non-linear optical performance of ZnO thin films. Appl. Phys. A. 2022;128:539. doi: 10.1007/s00339-022-05682-y. [DOI] [Google Scholar]
  • 26.Xiancai L., Xiaohua C., Wenjuan W., Yifeng Y., Guohua R. Preparation and characterization of WO3 from ammonium paratungstate via hydrothermal method. Front. Chem. 2006;1:389–392. doi: 10.1007/s11458-006-0057-2. [DOI] [Google Scholar]
  • 27.Pee J., Kim Y. Extraction factor of un-doped ammonium paratungstate from tungsten scraps. Arch. Metall. Mater. 2016;45:454–457 2016. doi: 10.1515/amm-2015-0141. [DOI] [Google Scholar]
  • 28.Salmaoui S., Sediri F., Gharbi N., Perruchot C., Aeiyach S., Rutkowska I.A., Kulesza P.J., Jouini M. Hexagonal nanorods of tungsten trioxide: synthesis, structure, electrochemical properties and activity as supporting material in electrocatalysis. Appl. Surf. Sci. 2011;257:8223–8229. doi: 10.1016/j.apsusc.2011.04.077. [DOI] [Google Scholar]
  • 29.Salmaoui S., Sediri F., Gharbi N., Perruchot C., Jouini M. Hexagonal hydrated tungsten oxide nanomaterials: hydrothermal synthesis and electrochemical properties. Electrochim. Acta. 2013;108:634–643. doi: 10.1016/j.electacta.2013.07.086. [DOI] [Google Scholar]
  • 30.Lokhande V., Lokhande A., Namkoong G., Kim J.H., Ji T. Charge storage in WO3 polymorphs and their application as super-capacitor electrode material. Results Phys. 2019;12:2012–2020. [Google Scholar]
  • 31.Das R., Sarkar S. Determination of intrinsic strain in poly (vinylpyrrolidone)-capped silver nano-hexapod using X-ray diffraction technique. Curr. Sci. 2015:775–778. https://www.jstor.org/stable/24905739 [Google Scholar]
  • 32.Langford J.I., Delhez R., De Keijser T.H.H., Mittemeijer E.J. Profile analysis for microcrystalline properties by the Fourier and other methods. Aust. J. Phys. 1988;41(2):173–188. doi: 10.1071/PH880173. [DOI] [Google Scholar]
  • 33.Warren B.E., Averbach B.L. The separation of cold-work distortion and particle size broadening in X-ray patterns. J. Appl. Phys. 1952;23:497. https://www.osti.gov/biblio/4378605 [Google Scholar]
  • 34.Balzar D., Ledbetter H. Voigt-function modeling in fourier analysis of size- and strain-broadened X-ray diffraction peaks. J. Appl. Crystallogr. 1993;26:97–103. doi: 10.1107/S0021889892008987. [DOI] [Google Scholar]
  • 35.Hall W.H. X-ray line broadening in metals. Proc. Phys. Soc. 1949;62:741–743. doi: 10.1088/0370-1298/62/11/110. [DOI] [Google Scholar]
  • 36.Tagliente M.A., Massaro M. Strain-driven (002) preferred orientation of ZnO nanoparticles in ion-implanted silica. Nucl. Instrum. Methods Phys. Res. B. 2008;266:1055. doi: 10.1016/j.nimb.2008.02.036. [DOI] [Google Scholar]
  • 37.Jacob R., Isac J. X-ray diffraction line profile analysis of Ba0.6Sr0.4FexTi(1-x)O3-δ, (x=0.4) Int. J. Chem. Stud. 2015;2:12–21. P-ISSN 2349–8528 E-ISSN 2321–4902. [Google Scholar]
  • 38.Dongale T.D., Mohite S.V., Bagade A.A., Kamat R.K., Rajun K.Y. Bio-mimicking the synaptic weights, analog memory, and forgetting effect using spray deposited WO3 memristor device. Microelectron. Eng. 2017;183:12–18. doi: 10.1016/j.mee.2017. [DOI] [Google Scholar]
  • 39.Au B.W.C., Chan K.Y., Knipp D. Effect of film thickness on electrochromic performance of sol-gel deposited tungsten oxide (WO3) Opt. Mater. 2019;94:387–392. doi: 10.1016/j.optmat.2019.05.051. [DOI] [Google Scholar]
  • 40.Goh E.G., Xu X., McCormick P.G. Effect of particle size on the UV absorbance of zinc oxide nanoparticles. Scripta Mater. 2014;78:49–52. doi: 10.1016/j.scripttanmat.2014.01.033. [DOI] [Google Scholar]
  • 41.Ghasemi L., Jafari H., Teacher S.R. Morphological characterization of tungsten trioxide nanopowders synthesized by sol-gel modified pechini's method leila. Mater. Res. 2017;20:1713–1721. doi: 10.1590/1980-5373-MR-2017-0467. [DOI] [Google Scholar]
  • 42.Kuznetsova M., Oliveira S.A., Rodrigues B.S., Souza J.S. Microwave - assisted synthesis of bismuth niobate/tungsten oxide photoanodes for water splitting. Top. Catal. 2020;3:345–349. doi: 10.1007/s11244-020-01325-9. [DOI] [Google Scholar]
  • 43.Sun Q.C., Ding Y., Goodman S.M., Funke H.H., Nagpal P. Copper plasmonics and catalysis: role of electron–phonon interactions in dephasing localized surface plasmons. Nanoscale. 2014;6:12450–12457. doi: 10.1039/C4NR04719B. [DOI] [PubMed] [Google Scholar]
  • 44.Taylor P., Davis E.A., Mott N.F. Conduction in non-crystalline systems V Conductivity, optical absorption and photoconductivity in amorphous semiconductors. Philos. Mag. A. 2006;22:37–41. doi: 10.1080/14786437008221061. [DOI] [Google Scholar]
  • 45.Abdelsalam D.G., Byung B.J., Abdel-Aziz F., Chegal W., Kim D. Highly accurate film thickness measurement based on automatic fringe analysis. Optik. 2012;123(16):1444–1449. doi: 10.1016/j.ijleo.2011.07.065. [DOI] [Google Scholar]
  • 46.You A., Be M.Y. I. In, Effect of silver incorporation in phase formation and band gap tuning of tungsten oxide thin films. J. Appl. Phys. 2012;112 doi: 10.1063/1.4768206. [DOI] [Google Scholar]
  • 47.Badawi A., Althobaiti M.G. Effect of Cu-doping on the structure, FT-IR and optical properties of Titania for environmental-friendly applications. Ceram. Int. 2021;47(8):11777–11785. doi: 10.1016/j.ceramint.2021.01.018. [DOI] [Google Scholar]
  • 48.Weinhardt L., Blum M., Bär M., Heske C., Cole B., Marsen B., Miller E.L. Electronic surface level positions of WO3 thin films for photoelectrochemical hydrogen production. J. Phys. Chem. C. 2008;112:3078–3082. doi: 10.1021/jp7100286. [DOI] [Google Scholar]
  • 49.Torad E., Ismail E.H., Mohamed M.M., Khalil M.M. Tuning the redox potential of Ag@ Ag2O/WO3 and Ag@ Ag2S/WO3 photocatalysts toward diclofenac oxidation and nitrophenol reduction. Mater. Res. Bull. 2021;137 doi: 10.1016/j.materresbull.2020.111193. [DOI] [Google Scholar]
  • 50.Jin J., Yu J., Guo D., Cui C., Ho W. A hierarchical Z‐scheme CdS–WO3 photocatalyst with enhanced CO2 reduction activity. Small. 2015;11:5262–5271. doi: 10.1002/smll.201500926. [DOI] [PubMed] [Google Scholar]
  • 51.Ghasemi L., Jafari H., Teacher S.R. Morphological characterization of tungsten trioxide nanopowders synthesized by sol-gel modified pechini's method leila. Mater. Res. 2017;20:1713–1721. doi: 10.1590/1980-5373-MR-2017-0467. [DOI] [Google Scholar]
  • 52.Raja M., Chandrasekaran J., Balaji M., Kathirvel P. Investigation of microstructural and optical and dc electrical properties of spin coated Al :WO3 thin films for n-Al : WO3/p-Si heterojunction diodes optik investigation of microstructural , optical and dc electrical properties of spin coated Al :WO3. Opt. - Int. J. Light Electron Opt. 2017;145:169–180. doi: 10.1016/j.ijleo.2017.07.049. [DOI] [Google Scholar]
  • 53.Marnadu R., Chandrasekaran J., Vivek P., Balasubramani V., Maruthamuthu S. Impact of phase transformation in WO3 thin films at higher temperature and its compelling interfacial role in Cu/WO3/p-Si structured Schottky barrier diodes,”. J. Phys. Chem. 2020;234:355–379. doi: 10.1515/zpch-2018-1289. [DOI] [Google Scholar]
  • 54.Aly S.A., Akl A.A., Mahmoud D.H. Microstructural and electrical characteristics of sprayed Tungsten oxide thin films. Int. J. New. Hor. Phys. 2015;52:47–52. doi: 10.12785/ijnhp/020202. [DOI] [Google Scholar]

Articles from Heliyon are provided here courtesy of Elsevier

RESOURCES